Ioannis
Mylonas-Margaritis
*a,
Zoi G.
Lada
b,
Alexandros A.
Kitos
a,
Diamantoula
Maniaki
a,
Katerina
Skordi
c,
Anastasios J.
Tasiopoulos
c,
Vlasoula
Bekiari
d,
Albert
Escuer
e,
Julia
Mayans
*e,
Vassilios
Nastopoulos
*a,
Evangelos G.
Bakalbassis
*f,
Dionissios
Papaioannou
*a and
Spyros P.
Perlepes
*ab
aDepartment of Chemistry, University of Patras, 26504 Patras, Greece. E-mail: ioanismylonasmargaritis@gmail.com; dapapaio@upatras.gr; nastopoulos@upatras.gr; perlepes@upatras.gr; bakalbas@chem.auth.gr
bInstitute of Chemical Engineering Sciences (ICE-HT), Foundation for Research and Technology-Hellas (FORTH), P.O. Box 1414, Platani, 26504 Patras, Greece
cDepartment of Chemistry, University of Cyprus, 1678 Nicosia, Cyprus
dDepartment of Agriculture, University of Patras, 26504 Patras, Greece
eDepartment de Quimica Inorganica i Organica, Secio Inorganica and Institute of Nanoscience (IN2UB) and Nanotechnology, Universitat de Barcelona, Marti i Franques 1-11, 08028 Barcelona, Spain. E-mail: julia.mayans@qi.ub.edu
fDepartment of Chemistry, Aristotle University of Thessaloniki, University Campus, 54124 Thessaloniki, Greece. E-mail: bakalbas@chem.auth.gr
First published on 15th May 2023
The initial use of a tetradentate Schiff base (LH2) derived from the 2:
1 condensation between 2-hydroxyacetophenone and cyclohexane-1,2-diamine in 4f-metal chemistry is described. The 1
:
2 reaction of Ln(NO3)3·xH2O (Ln = lanthanoid or yttrium) and LH2 in MeOH/CH2Cl2 has provided access to isostructural complexes [Ln(NO3)3(L′H2)(MeOH)] in moderate to good yields. Surprisingly, the products contain the corresponding Schiff base ligand L′H2 possessing six aliphatic –CH2– groups instead of the –CH–(CH2)4–CH– unit of the cyclohexane ring, i.e. an unusual ring-opening of the latter has occurred. A mechanism for this LnIII-assisted/promoted LH2 → L′H2 transformation has been proposed assuming transient LnII species and a second LH2 molecule as the H2 source for the reduction of the cyclohexane moiety. DFT calculations provide strong evidence for the great thermodynamic stability of the products in comparison with analogous complexes containing the original intact ligand. The structures of the PrIII, SmIII, GdIII, TbIII, and HoIII complexes have been determined by single-crystal X-ray crystallography. The 9-coordinate LnIII centre in the molecules is bound to six oxygen atoms from the three bidentate chelating nitrato groups, two oxygen atoms that belong to the bidentate chelating organic ligand, and one oxygen atom from the coordinated MeOH group. In the overall neutral bis(zwitterionic) L′H2 ligand, the acidic H atoms are clearly located on the imino nitrogen atoms and this results in the formation of an unusual 16-membered chelating ring. The coordination polyhedra defined by the nine donor atoms around the 4f-metal-ion centres can be best described as distorted, spherical capped square antiprisms. The EuIII, TbIII, and DyIII complexes exhibit LnIII-based luminescence in the visible region, with the coordinated L′H2 molecule acting as the antenna. Ac magnetometry experiments show that the DyIII member of the family behaves as an SIM at zero field and under external dc fields of 0.1 and 0.2 T without the enhancement of the peaks’ maxima, suggesting that QTM is not the relaxation path. The GdIII complex behaves, rather unexpectedly, as a SIM with two different magnetic relaxation paths occurring at very close temperatures; this behaviour is tentatively attributed to a very small axial zero-field splitting (D ∼ 0.1 cm−1), which cannot be detected by magnetization or susceptibility experiments. The prospects of the present, first results in the lanthanoid(III)-LH2 chemistry are discussed.
As far as the magnetic properties of molecular LnIII complexes are concerned, the discovery of slow relaxation of the magnetization and magnetic hysteresis in a bis(phthalocyaninato)terbium(III) complex in 20035 ignited an explosive growth of research interest in LnIII-based molecular nanomagnets, either coordination clusters (single-molecule magnets, SMMs) or mononuclear complexes (single-ion magnets, SIMs), because of their exciting physical phenomena and potential applications in magnetic memory storage, molecular spintronics, and quantum computing.6,7 Design principles developed by Rinehart and Long directed synthetic chemists and physicists towards longer relaxation times by means of extremely large increases in the energy barrier to magnetization reversal (Ueff).8 The massive increases did not lead to corresponding increases in the blocking temperature (TB) due to the vibronically-induced spin relaxation mechanisms – irrespective of the Ueff value – that can operate.9 A breakthrough happened in 2017 when it was shown10 that the cation [Dy(Cpttt)2]+, where Cpttt is C5H5tBu-1,2,4, exhibits magnetic hysteresis up to 60 K. Subsequent studies showed impressive results in this area.11a,b Removal of C–H groups from the C5 ring of Cpttt through the synthesis of peralkylated bis(cyclopentadienyl) DyIII complexes led to hysteresis temperatures TB around the ligand nitrogen temperature (∼80 K), which is the current record.11c
As far as the optical properties are concerned, most LnIII ions luminesce in the solid state. Unlike luminescence from organic compounds, most 4f-metal ion emissions consist of sharp lines. This property has been used in EuIII and TbIII phosphors, and in the NdIII YAG laser. EuIII and TbIII, and sometimes DyIII, display luminescence in the visible region.12 The 4f → 4f transitions are Laporte forbidden and thus excitation of LnIII to an emissive state by this route is not an efficient process. An alternative method of excitation is via an organic ligand, usually an aromatic system, which has an excited triplet state higher in energy than the LnIII emissive state, the so-named antenna effect.13
We are interested in mononuclear LnIII complexes that show magnetic relaxation (SIMs) and/or LnIII-based luminescence. For the realization of this general goal, the choice of the primary organic ligand is very important. For the preparation of SIMs, the ligand should behave as terminal (either monodentate or chelating). It is also crucial for the appearance of the required magnetic anisotropy of the molecule and a high separation between Mj and Mj ± 1 states (in order to obtain high Ueff values), and these imply a rational design of the ligand field. For oblate-type ions (e.g. TbIII and DyIII), the ligand donor atoms with the greatest electron density should bind at axial positions enhancing the required axial anisotropy, whereas for prolate-type ions (e.g. ErIII and YbIII), the ligand donor atoms with the largest electron density should coordinate to equatorial positions to achieve axial anisotropy. For the synthesis of luminescent LnIII complexes with metal ion-based emission, the ligands should bear chromophores which can facilitate efficient energy transfer “sensitization” of the LnIII ion's excited (i.e. emissive) levels from the ligand's triplet state. Schiff bases are often used for these purposes.14 We have been using Schiff bases of various types to achieve the above-mentioned objectives.15
We would also like to mention that the chemistry of mononuclear LnIII-Schiff base complexes is attracting the interest of inorganic chemists for two additional reasons. First, it has been shown that two and four electrons can be stored, respectively, in intramolecular and intermolecular C–C bonds formed by LnIII-assisted reduction of the imino group of Schiff-base ligands, providing new synthetic avenues to the reductive chemistry of lanthanoids.16 Second, complexes of YbIII with chelating Schiff bases are qubit candidates because of the huge splitting between the electronic ground doublet and the first excited crystal field state and their intrinsic slow paramagnetic relaxation,17 as well as candidates for novel coupled electronic qubit-nuclear qubit systems.18
We have recently embarked on a new subarea of LnIII-Schiff base chemistry by using N,N′-bis(salicylidene)ethylenediamine (salen)-type ligands, in which non-conjugated, e.g. cyclohexyl, bridges are linked with two salicylaldehyde moieties functionalised either on the aromatic rings and/or at the aliphatic carbon atoms. Our general goal is to compare the magnetic (e.g. SIM responses) and optical (e.g. aggregation-induced emission, AIE, or aggregation-caused quenching, ACQ, behaviour19) properties of the LnIII complexes containing functionalized Schiff bases with those of the corresponding complexes containing non-functionalized ligands. We plan to use various electron-accepting (–NO2, –F, –Cl, –CN, –Ph), electron-donating (–OMe, –OH, –NEt2, –Me) or bulky (e.g.tBu) substituents. This work started with the reactions of lanthanoid(III) nitrates with the Schiff base LH2 (Scheme 1), which possesses methyl groups at the aliphatic carbon atoms of the salicylaldehyde moiety. No LnIII complexes of the neutral or anionic forms of LH2 have been reported; in contrast, LnIII complexes of salcnH2 or salcn2−, i.e. the unsubstituted Schiff base, have been synthesized and structurally characterised, and their emission and magnetic properties were studied.20 However, we were surprised to discover that LH2 undergoes a LnIII-promoted/assisted cyclohexane ring opening and the isolated complexes contain a neutral L′H2 ligand (Scheme 1). The interesting structural features and properties (optical, magnetic) of the products are described in this article.
Analytical data, calcd for C23H32LnN5O12 (found values are in parentheses): 2: C 38.64 (39.13), H 4.52 (4.47), N 9.80 (9.66); 3: C 38.31 (37.99), H 4.48 (4.59), N 9.72 (9.63); 4: C 38.24 (38.57), H 4.47 (4.56), N 9.69 (9.75); 5: C 37.95 (37.43), H 4.47 (4.41), N 9.62 (9.39); 6: C 37.86 (37.98), H 4.43 (4.52), N 9.60 (9.41); 7: C 37.68 (37.92), H 4.41 (4.32), N 9.56 (9.48); 8: C 37.56 (38.09), H 4.39 (4.28), N 9.52 (9.40); 9: C 37.44 (37.80), H 4.38 (4.49), N 9.49 (9.52); 10: C 37.15 (36.99), H 4.35 (4.38), N 9.42 (9.29); 11: C 41.88 (42.04), H 4.90 (4.81), N 10.62 (10.55)%.
Important crystallographic data are listed in Table S1.† Full details can be found in the CIF files. Crystallographic data for the structures reported in this paper have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication with the deposition numbers 2246267–2246271.†
Since the observed transformation was totally unexpected, we re-examined carefully the purity of LH2. Also we examined the purity of the 1,2-diaminocyclohexane precursor to investigate if it contained a portion of 1,6-diaminohexane whose condensation with 2-hydroxyacetophenone would have given L′H2 as a contaminant of LH2; in the case of contamination, the L′H2 component of the mixture with LH2 would result in the isolation of 1–11. Our studies proved the purity of LH2 and its 1,2-diaminocyclohexane precursor. Thus, (1) LH2 was obtained in the form of yellow single crystals; the unit cell determination of a few of them revealed that they represent21 the structurally characterised compound (E,E)-2,2′-[1,1′-(cyclohexane-1,2-diyldinitrilo)diethylidene]diphenol (Scheme 2); for the preparation of the complex, we used batches of single crystals which were subsequently powdered. (2) The 1H NMR spectrum of the used powder in d6-DMSO shows the expected32 signals for pure LH2 with the correct integration ratio; there are four signals in the aromatic region (δ 7.54–6.83 ppm) representing eight protons, a doublet of doublet signal at δ 4.02 ppm due to the two protons attached to positions 1 and 2 of the cyclohexane ring, a singlet signal at δ 2.38 ppm representing the six protons of the two methyl groups and a group of three multiplet signals at δ values of 2.07 (four protons), 1.84 (two protons) and 1.64 (two protons) ppm which represent the eight –CH2– protons of the cyclohexane ring. The broad singlet at δ ∼13.5 ppm (two protons) is assigned19 to the –OH hydrogen atoms. (3) The 1H NMR spectrum of (±)-trans-1,2-diaminocyclohexane (used for the synthesis of LH2) in d6-DMSO confirms its purity. The spectrum (Fig. S1†) shows five multiplet signals at δ 2.03, 1.70, 1.58, 1.17 and 0.96 ppm with an integration ratio of 1:
1
:
1
:
1
:
1 (or 2
:
2
:
2
:
2
:
2). The signal at a lower field (δ 2.03 ppm) represents the two protons at the 1 and 2 positions of the cyclohexane ring (i.e. the protons belonging to the carbon atoms attached to the two nitrogen atoms), while the other signals correspond to the eight –CH2– protons. The –NH2 protons do not appear because of their exchangeable character. The spectrum is identical (as expected) to the spectra of the pure (+)-S,S-trans (Fig. S2†) and (−)-R,R-trans enantiomers of 1,2-diaminocyclohexane (provided by Fluorochem). Moreover, this spectrum is different from the recorded spectrum of 1,6-diaminohexane in d6-DMSO; the latter shows a triple signal at δ 2.68, a multiplet at δ 1.35 and a multiplet at δ 1.12 with an integration ratio of 1
:
1
:
1 (or 4
:
4
:
4) attributed to the twelve –(CH2)– protons. (4) The 13C NMR spectrum of (±)-trans-1,2-diaminocyclohexane, in d6-DMSO (Fig. S3†), which is again identical to the 13C NMR spectra of the pure (+)-S,S-trans and (−)-R,R-trans, (Fig. S4†) enantiomers, shows the three expected signals of the cyclohexane ring at δ values of 58.1, 35.4 and 25.7 ppm, the former being assigned to the carbons attached to the nitrogen atoms. Moreover, this spectrum is different from the recorded 13C spectrum of 1,6-diaminohexane, which displays the three expected signals at δ values of 41.8 (assigned to the C atoms attached to nitrogen), 32.8 (assigned to the C atoms at the 2 and 5 positions) and 26.8 (assigned to the C atoms at the 3 and 4 positions). All the above results indicate that our starting material (±)-trans-1,2-diaminocyclohexane is pure, and not contaminated with 1,6-diaminohexane.
![]() | ||
Scheme 2 Schematic illustration of the E,E configuration of LH2 as observed in its single-crystal X-ray structure.21 |
The molar conductivity (ΛM) values of 10−3 M solutions of 1–11 are in the range of 85–110 S cm2 mol−1 indicating 1:
3 electrolytes33 and the decomposition of the complexes in solution.
![]() | ||
Fig. 2 The spherical capped square antiprismatic coordination polyhedron of PrIII in complex 1. The very small spheres define the vertices of the ideal polyhedron. |
The crystal structure of 1 consists of [Pr(NO3)3(L′H2)(MeOH)] molecules. The 9-coordinate PrIII centre is bound to six oxygen atoms from the three slightly anisobidentate chelating nitrato groups, two oxygen atoms that belong to the bidentate chelating L′H2 ligand and one oxygen atom from the coordinated MeOH molecule. The acidic H atoms, which are bound to the oxygen atoms of the hydroxyl groups in the structure of the free ligand,21 are clearly located on the imino nitrogen atoms (N1, N2); thus, the formally neutral ligand participates in its bis(zwitterionic) form leaving the oxygen atoms deprotonated, blocking the N coordination sites and forming an unusual 16-membered chelating ring. The two NC(imino)–C(aromatic)–C(aromatic)–O parts of the ligand and the PrIII centre are almost coplanar, with the largest deviation from their best mean plane being that of O2 (0.137(2) Å); the two aromatic rings are also coplanar and the angle between them is 1.2(2)°. The PrIII–O bond lengths fall in the range of 2.334(2)–2.616(2) Å and are typical of 9-coordinate PrIII complexes. The bond lengths of PrIII to deprotonated phenolato oxygens (2.334(2) and 2.337(2) Å) are shorter than the distances of the nitrato groups (2.534(2)–2.616(2) Å) and methanol (2.501(2) Å). There are two strong intramolecular H bonds in the complex molecule (Table S2†) with the protonated nitrogen atoms being the donors and the negatively charged coordinated oxygen atoms being the acceptors.
There are no Platonic, Archimedean and Catalan polyhedra with nine vertices, and also this number of vertices cannot result in prisms or antiprisms. Thus, the only shapes that may be considered are those listed in Table S3.† With the help of program SHAPE,34 the best fit obtained for the PrIII coordination polyhedron is for the spherical capped square antiprism, with the nitrato atom O9 being the capping atom. Since the coordinated nitrato groups impose small coordination angles (∼50°), the polyhedron is distorted (Fig. 2).
A variety of intermolecular interactions stabilize the crystal structure of 1. The two molecules of the unit cell form a centrosymmetric dimer bridged by two relatively strong (Table S2†) O12–H12⋯O11 H bonds (Fig. 3) about the centre of symmetry located at ½, ½, ½ of the cell. The molecules are further packed in a 3D network through weak C–H⋯O interactions.
Each analogous Ln–O bond length in the five isomorphous complexes follows the order Pr > Sm > Gd > Tb > Ho, which is a consequence of the LnIII contraction.
Compounds 1, 3, 5, 6 and 8 are the first structurally characterized complexes of L′H2 or/and its anionic form with any metal. Homometallic LnIII complexes of the salicylaldehyde analogues of L′H2, i.e. salhexnH235 (Scheme 1), have been reported.36 They are: (i) {[Eu2(hfac)4(O2CMe)2(salhexnH2)2]}n,36a a 1D coordination polymer in which the neutral Schiff-base molecules behave as O,O′-bidentate bridging ligands with the acidic H atoms being located at the oxygen atoms (hfac− is the hexafluoroacetylacetato ligand); (ii) [Eu2(dbm)4(O2CMe)2(salhexnH2)2],36a where the two EuIII centres are linked by two O,O′-bidentate bridging ligands with the location of the acidic H atoms being not reported (dbm− is the dibenzoylmethanide ligand); (iii) {[Gd2(tta)4(O2CMe)2(salhexnH2)2]}n,36b a 1D polymer in which the neutral salhexnH2 molecules behave as O,O′-bidentate bridging ligands with the acidic H atoms being located at the imino nitrogen atoms (tta− is the 2-thenoyltrifluoroacetylacetonato ligand). Analogous O,O′-bidentate bridging coordination modes have been reported for homometallic LnIII polymers (1D, 2D)37,38 and dimers36a possessing the –(CH2)4– (i.e. N,N′-bis(salicylidene-1,4-butanediamine)36a,37 and –(CH2)5–38 (i.e. N,N′-bis(salicylidene-1,5-pentanediamine)38 analogues of salhexnH2. Thus, the O,O′-bidentate chelating mode of L′H2 in 1–11 is novel in the coordination chemistry of such ligands with aliphatic backbones containing four or more carbon atoms, not only with lanthanoids but also with other metal ions. All the above complexes of the salhexnH2, and –(CH2)4– and –(CH2)5– analogues were prepared using the open-chain Schiff base which was incorporated in the complexes; no transformation similar to the LH2 → L′H2 one reported in this work was noticed.
![]() | ||
Fig. 5 Solid-state, room-temperature excitation (curve 1; maximum emission at 616 nm) and emission (curve 2; maximum excitation at 399 nm) spectra of solid [Eu(NO3)3(L′H2)(MeOH)] (4). |
![]() | ||
Fig. 6 Solid-state, room-temperature excitation (curve 1; maximum emission at 488 nm) and emission (curve 2; maximum excitation at 405 nm) spectra of solid [Tb (NO3)3(L′H2)(MeOH)] (6). |
The solid-state, room-temperature luminescence spectra of the EuIII (4), TbIII (6) and DyIII (7) complexes, which are candidates for LnIII-based emission in the visible region,12,13 were recorded. Upon maximum excitation at 399 nm, a solid sample of 4 displays photoluminescence with maxima at 595, 616, 652 and 695 nm due to EuIII.13a,b,42 These peaks are assigned to the characteristic 5D0 → 7Fj (j = 0–4) transitions. Specific assignments are as follows: 5D0 → 7F1,2 (595 nm), 5D0 → 7F2 (616 nm), 5D0 → 7F3 (652 nm) and 5D0 → 7F4 (695 nm). The dominant peak at 616 nm is due to the hypersensitive 5D0 → 7F2 transition. The higher intensity of this transition compared with that of the magnetic-dipole allowed 5D0 → 7F1 transition indicates that this complex has no imposed symmetry at EuIII, in accordance with our belief that 4 is isomorphous with the structurally characterized complexes 1, 3, 5, 6 and 8.
Upon maximum excitation at 405 nm, solid 6 displays photoluminescence with maxima at 488, 543, 588, 621 (shoulder) and 648 nm due to TbIII.42c,43 These emission maxima are assigned to the transitions from the 5D4 state to its 7F6 (488 nm), 7F5 (543 nm), 7F4 (588 nm), 7F3 (621 nm) and 7F2 (648 nm) states.
The excitation spectrum of complex 7 (Fig. S12†) shows medium-to-strong intensity bands which are attributed44 to the appropriate transitions of DyIII: 6H15/2 → 4M17/2, 6P7/2 (330 nm), 6H15/2 → 2K17/2 (380 nm) and 6H15/2 → 4I15/2 (418 nm). Upon maximum excitation at 380 nm, solid 7 displays photoluminescence with maxima at 486 and 519 nm being attributed to the intraconfigurational 4f–4f transitions of the metal ion (Fig. S12†).44 The maximum located at 486 nm can be safely assigned to the magnetic dipole transition 4F9/2 → 6H15/2, which generally is not sensitive to the crystal field. The expected “yellow” 4F9/2 → 6H13/2 around 575 nm is not observed in our spectrum; this emission is related to the forced electric dipole transition-type and its intensity is strongly influenced (and often strongly quenched) by the crystal-filed environment. Thus, the origin of the 519 nm maximum is not clear.
Upon cooling, the value of the χMT product for the isotropic (4f7) GdIII complex 5 is constant down to 25 K and then decreases slowly to a value of 6.0 cm3 K mol−1 at 2.0 K. The low-temperature decrease of the product should be attributed to weak intermolecular antiferromagnetic exchange interactions promoted by the two intermolecular Omethanol–H⋯“free” Onitrato H bonds (vide supra, Table S2†) which form {GdIII}2 dimers. The fit of the experimental data, applying the spin Hamiltonian H = −2J(S1·S2), give a satisfactory simulation for J = −0.03 cm−1 with a g value of 1.98. Magnetization measurements at 2.0 K show a saturation value of 7.03Nμβ in agreement with the expected 7/2 spin. Reduced magnetization data indicate a quasi-negligible anisotropy (Fig. 8, right).
For the TbIII (6), DyIII (7) and HoIII (8) complexes, the values of the χMT product decrease slowly when lowering T due to the progressive depopulation of the mj levels of the ground state J, with a more pronounced decay below 50 K; the final values at 2.0 K are 7.01 (6), 9.21 (7) and 5.07 (8) cm3 K mol−1. The sharp decay at low temperatures might also be due to weak antiferromagnetic LnIII⋯LnIII interactions within the H-bonded dimers. The susceptibility data were fitted assuming implicitly a regular distribution of the mj states. The Hamiltonian used is represented by eqn (1), where Ŝ is the spin operator, L is the orbit operator,45,46λ is the spin–orbit coupling and Δ describes the energy gap between mL components. The value for the orbital reduction parameter k was assumed as 1 (the LnIII ions behave as purely ionic). Thus, the first term of the Hamiltonian describes the spin–orbit coupling, the second is related to the ligand field around the lanthanoid cation, and the third term describes the Zeeman effect.
H = (λŜL) + Δ[LZ2 − L(L + 1)/3] + [βH(−kL + 2Ŝ)] | (1) |
In spite of the low symmetry of the LnIII environment in 6–8, the χMT simulations were accurate with Δ parameters of −37, −10 and −14 cm−1 for 6, 7 and 8, respectively. The negative values of Δ indicate that the highest mj value corresponds to the ground state. The magnetization plots of 6–8 at 2.0 K show a fast increase at a low field with quasi saturation (Fig. 8), confirming the negative Δ values for these complexes.
Ac magnetic susceptibility measurements were performed using a 4.0 G ac field on the powdered samples of 6–8 under different static fields in order to explore the magnetization dynamics of the three complexes. No out-of-phase (imaginary) components of the ac susceptibility, χ′′M, were detected in the TbIII and HoIII complexes at frequencies between 10 and 1488 Hz, either at zero field or under external applied fields. In contrast, DyIII complex 7 exhibits tails of peaks below 5 K as expected for a Kramers’ ion with mj = ±15/2 ground doublet (Fig. 9, left and S15†). The χ′′M response was measured at zero dc field, and also under applied external (bias) dc fields of 0.1 and 0.2 T without observing any enhancement of the peaks’ maxima; this indicates that the quantum tunnelling of the magnetization (QTM) is not the relaxation path in this case despite the distorted environment around the DyIII centre. Due to the lack of maxima in the 10–1480 Hz frequency range and in the measured temperature range, the system was fitted (Fig. 9, right and S14†) according to the Debye eqn (2),47 giving the mean values of 4.6 cm−1 and 7.6 × 10−7 s for the effective barrier for the magnetization reversal (Ueff) and pre-experimental factor (τ0), respectively. The experimental data were not enough for a complete Cole–Cole plot to investigate the magnetic relaxation mechanism; however, due to the Ueff value which is smaller than the energy of the first excited state, the Orbach relaxation is discarded as the main relaxation path.
ln(χ′′M/χ′M) = ln(ωτ0) − Ueff/kBT | (2) |
Given the recent discovery of and great interest on GdIII SIMs/SMMs,48,49 we investigated the dynamic properties of 5. Out-of-phase magnetic susceptibility signals were not observed at zero field, but upon increasing the static (dc) field tails of signals appeared with a maximum intensity around 0.4 T. The χ′′Mversus T measurements were carried out at this field in the 10–1488 Hz range and a clear frequency dependence was observed, Fig. 10. The χ′′M(T) plot shows two processes, one as a shoulder signal at higher frequencies around 2.5 K and a second one below 1.8 K, which are almost completely overlapped showing that the two different magnetic relaxation paths are occurring at very close temperatures (Fig. 10, top). A fit of the Argand plot performed by the generalized Debye model shows an α parameter with values between 0.06 and 0.16 indicating a narrow distribution of relaxation times and short τ values in the 10−5–10−6 s interval which are shown as ln(τ) vs. the inverse of temperature in Fig. 10, bottom and S14.† Fitting of the ln(τ) plot can be performed with similar quality factors combining several relaxation mechanisms such as Orbach plus Raman, Orbach plus direct or direct plus Raman. However, taking into account that the conventional Ueff barrier derived from a double-well potential cannot be applied for this kind of quasi-isotropic system (DS2 ≪ kT), the only realistic fit is from the direct (low temperature) plus Raman (high temperature, via lattice vibrations) relaxation combination applying the expression:
τ−1 = AT + CTn | (3) |
![]() | ||
Fig. 10 Top, temperature dependence (a) and frequency dependence (b) of χ′′M for complex 5 under a 0.4 T static field. Bottom, Argand plot (c) and fit of the ln(τ) vs. inverse of temperature (d). |
The best fit values were A = 9136, C = 6.85 and n = 4.90, Fig. 10 bottom. The n value is lower than the theoretical n = 9 value for the Kramers ions but is commonly attributed to the participation of optical lattice vibrations, which has been previously reported.50
Complex 5 joins a limited number of GdIII complexes exhibiting slow relaxation of the magnetization in spite of the isotropic character of the cation, which contrasts with the classical interpretation for the slow relaxation of the magnetization usually explained as an over barrier process depending on the axial zero field splitting. The mostly accepted explanation48a,g,h,49 is the presence of a weak axial zero-field splitting (D ∼ 0.1 cm−1), which cannot be detected by magnetization or susceptibility studies. These small values, which result from molecular distortion, allow very small double-well energy barriers around ∼1 cm−1 (calculated by the formula DS2 − 1/4), which are in the orders of magnitude lower than the experimentally calculated barriers (Ueff). Such low D values, however, imply that the relaxation phenomenon cannot be explained in terms of conventional over-barriers or QTM above 2 K. The weak anisotropy breaks the degeneration at the S = 7/2 level, mixing the mS sublevels under moderate (∼0.4 T) external magnetic fields and resulting in SIM/SMM behaviour.
Our future directions in the project include: (a) Reactions of LH2 with 3d and 5f (uranyl, thorium(IV)) ions to see if the LH2 → L′H2 transformation can be observed again; preliminary experiments indicate that 3d-metal complexes containing either LH2 or L′H2 can be isolated. For example, compounds {FeIIICl3(LH2)·Et2O}n, {[FeIIICl3(LH2)·2Me2CO}n and {[MnIICl2(LH2)2]·3MeOH}n have been recently structurally characterized, depending on the reaction conditions. Complex (Et3NH)[MnIII2(L)2(L′H2)](ClO4)3·CHCl3, which contains both the original and transformed ligands, can be isolated from the 1:
1
:
2 Mn(ClO4)2·6H2O/LH2/Et3N reaction mixture in CHCl3–MeOH (Fig. S16–S19†) (b) Synthesis and full characterization of L′H2 (this compound is not known), followed by its reaction with lanthanoid(III) nitrates to investigate if the same complexes (1–11) with those described here can be obtained. (c) Reactions of ligands analogous to LH2, but possessing various substituents (–Cl, –Br, –Me,⋯) on the aromatic rings, with LnIII ions to investigate if the identity of the products depends on the nature of the substituent and if similar metal ion-assisted transformations are operative; (d) reactions of Schiff bases derived from the 2
:
1 condensation between 2-hydroxyacetophenone and 1,2-diaminocyclopentane or 1,2-diaminocyclobutane with LnIII ions; the ring-opening process might be easier with such Schiff bases given the larger strain of these rings than that in cyclohexane (cyclobutane 27 kcal mol−1, cyclopentane 5 kcal mol−1, and cyclohexane 1 kcal mol−1).
Footnote |
† Electronic supplementary information (ESI) available: Crystallographic (Table S1) and structural data (Tables S2 and S13), various spectra (Fig. S1–S4, S10–S12), structural plots (Fig. S5–S9 and S16–S19), magnetic plots (Fig. S13–S15), mechanistic schemes (Schemes S1–S6), and computational drawings (Schemes S7 and S8). CCDC 2246267–2246271. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3dt00817g |
This journal is © The Royal Society of Chemistry 2023 |